15
MARINE ECOLOGY PROGRESS SERIES Mar Ecol Prog Ser Vol. 483: 117–131, 2013 doi: 10.3354/meps10261 Published May 30 INTRODUCTION Projections of global climate change forecast rapid atmospheric and oceanic temperature increases over the next 100 yr (IPCC 2007), which will cause changes in the strength and pattern of winds and ocean currents (Cai et al. 2005, Poloczanska et al. 2007). Consequently, population declines and range shifts of species are predicted to occur, with ramifica- tions for marine ecosystems (Sagarin et al. 1999, Parmesan 2006, Wernberg et al. 2011). Empirical studies support such predictions of climate impacts: changes in species phenology and distribution can alter community dynamics through flow-on effects to abiotic processes and food webs (Parmesan 2006, Holbrook et al. 2009, Beaugrand et al. 2002). Under- © Inter-Research 2013 · www.int-res.com *Email: [email protected] Variation in the morphology, reproduction and development of the habitat-forming kelp Ecklonia radiata with changing temperature and nutrients Christopher J. T. Mabin 1, *, Paul E. Gribben 2 , Andrew Fischer 1 , Jeffrey T. Wright 1 1 National Centre for Marine Conservation and Resource Sustainability (NCMCRS), Australian Maritime College, University of Tasmania, Launceston, Tasmania 7250, Australia 2 Biodiversity Research Group, Climate Change Cluster, University of Technology Sydney, Sydney, New South Wales 2007, Australia ABSTRACT: Increasing ocean temperatures are a threat to kelp forests in several regions of the world. In this study, we examined how changes in ocean temperature and associated nitrate concentrations driven by the strengthening of the East Australian Current (EAC) will influence the morphology, reproduction and development of the widespread kelp Ecklonia radiata in south- eastern Australia. E. radiata morphology and reproduction were examined at sites in New South Wales (NSW) and Tasmania, where sea surface temperature differs by ~5°C, and a laboratory experiment was conducted to test the interactive effects of temperature and nutrients on E. radiata development. E. radiata size and amount of reproductive tissue were generally greater in the cooler waters of Tasmania compared to NSW. Importantly, one morphological trait (lamina length) was a strong predictor of the amount of reproductive tissue, suggesting that morphological changes in response to increased temperature may influence reproductive capacity in E. radiata. Growth of gametophytes was optimum between 15 and 22°C and decreased by > 50% above 22°C. Microscopic sporophytes were also largest between 15 and 22°C, but no sporophytes developed above 22°C, highlighting a potentially critical upper temperature threshold for E. radiata in Tasmania. Lower nitrate concentration had no effect on E. radiata gametophytes and sporophytes. Given forecast increases in ocean temperature of between 2 and 3°C in southeastern Australia by 2100, these findings suggest that E. radiata is likely to be affected by a strengthening EAC and highlight the susceptibility of the development and growth of early life-cycle stages to these changes. KEY WORDS: Ecklonia radiata · East Australian Current · Climate change · Morphology · Reproduction Resale or republication not permitted without written consent of the publisher

Variation in the morphology, reproduction and development

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Marine Ecology Progress Series 483:117Vol. 483: 117–131, 2013 doi: 10.3354/meps10261
Published May 30
INTRODUCTION
Projections of global climate change forecast rapid atmospheric and oceanic temperature increases over the next 100 yr (IPCC 2007), which will cause changes in the strength and pattern of winds and ocean currents (Cai et al. 2005, Poloczanska et al. 2007). Consequently, population declines and range
shifts of species are predicted to occur, with ramifica- tions for marine ecosystems (Sagarin et al. 1999, Parmesan 2006, Wernberg et al. 2011). Empirical studies support such predictions of climate impacts: changes in species phenology and distribution can alter community dynamics through flow-on effects to abiotic processes and food webs (Parmesan 2006, Holbrook et al. 2009, Beaugrand et al. 2002). Under-
© Inter-Research 2013 · www.int-res.com*Email: [email protected]
Variation in the morphology, reproduction and development of the habitat-forming kelp Ecklonia radiata with changing temperature and nutrients
Christopher J. T. Mabin1,*, Paul E. Gribben2, Andrew Fischer1, Jeffrey T. Wright1
1National Centre for Marine Conservation and Resource Sustainability (NCMCRS), Australian Maritime College, University of Tasmania, Launceston, Tasmania 7250, Australia
2Biodiversity Research Group, Climate Change Cluster, University of Technology Sydney, Sydney, New South Wales 2007, Australia
ABSTRACT: Increasing ocean temperatures are a threat to kelp forests in several regions of the world. In this study, we examined how changes in ocean temperature and associated nitrate concentrations driven by the strengthening of the East Australian Current (EAC) will influence the morphology, reproduction and development of the widespread kelp Ecklonia radiata in south- eastern Australia. E. radiata morphology and reproduction were examined at sites in New South Wales (NSW) and Tasmania, where sea surface temperature differs by ~5°C, and a laboratory experiment was conducted to test the interactive effects of temperature and nutrients on E. radiata development. E. radiata size and amount of reproductive tissue were generally greater in the cooler waters of Tasmania compared to NSW. Importantly, one morphological trait (lamina length) was a strong predictor of the amount of reproductive tissue, suggesting that morphological changes in response to increased temperature may influence reproductive capacity in E. radiata. Growth of gametophytes was optimum between 15 and 22°C and decreased by >50% above 22°C. Microscopic sporophytes were also largest between 15 and 22°C, but no sporophytes developed above 22°C, highlighting a potentially critical upper temperature threshold for E. radiata in Tasmania. Lower nitrate concentration had no effect on E. radiata gametophytes and sporophytes. Given forecast increases in ocean temperature of between 2 and 3°C in southeastern Australia by 2100, these findings suggest that E. radiata is likely to be affected by a strengthening EAC and highlight the susceptibility of the development and growth of early life-cycle stages to these changes.
KEY WORDS: Ecklonia radiata · East Australian Current · Climate change · Morphology · Reproduction
Resale or republication not permitted without written consent of the publisher
Mar Ecol Prog Ser 483: 117–131, 2013
standing which life-cycle stages are most vulnerable to oceanographic changes and the mechanisms underpinning vulnerability (e.g. juveniles or adults) are critical to predicting the response of species to climate change (Russell et al. 2011).
In southeastern Australian coastal waters, the East Australian Current (EAC) is the dominant mode of oceanic input, particularly during summer (Ridgway 2007). EAC-associated waters are warmer and nitrate poor compared to the Antarctic-influenced waters that otherwise dominate the region, and recent strengthening of the EAC means warm and nitrate-poor waters are pushing farther south for longer periods of time (Rochford 1984, Cai et al. 2005, Cai 2006, Ridgway 2007). For example, long-term sea surface temperature (SST) datasets taken at the southern edge of the EAC (Maria Island, Tasmania) show an increase in annual average SST of >1°C since 1944 (Johnson et al. 2011). Additionally, at both Maria Island and a site that is approximately 8.5° lat- itude further north in New South Wales (NSW) (Port Hacking), SST is weakly negatively correlated with nitrate concentrations (Rochford 1984). Significantly, modeling predicts future extension of the February mean 20°C SST isotherm into currently cool-temper- ate Tasmanian waters, exposing that region to warm- temperate conditions by the end of the century (Bartsch et al. 2012). Overall, Tasmania’s east coast marine communities are predicted to be exposed to increasing SSTs ~3.8 times the global average (2.0 to 3.0°C) and associated longer exposure times to low nitrate levels (i.e. nitrates <1 µmol l−1) (Rochford 1984, Ridgway 2007).
Some major ecological effects of the strengthening of the EAC are already evident. These include the range expansion of the barrens-forming sea urchin Centrostephanus rodgersii, which is impacting sub - tidal kelp forests in Tasmania (Johnson et al. 2005, Ling et al. 2009a,b), and range shifts in a number of macroalgae (Wernberg et al. 2011) including the con- traction and reduction in canopy cover of giant kelp Macrocystis pyrifera in Tasmania (Johnson et al. 2011). The effect of a strengthening EAC on habitat- forming temperate kelp and associated species is of major importance. Globally, kelp facilitate high lev- els of temperate marine biodiversity (Anderson et al. 1996, Smith et al. 1996, Wernberg et al. 2009), and in southern Australia, kelp habitat (parti cularly Ecklo- nia radiata) contributes to the world’s highest rates of regional endemism in both marine algae and inverte- brates (Bolton 1996, Phillips 2001, Kerswell 2006). In Tasmania, the economic importance of kelp is recog- nised through support to fisheries-based ecosystems,
particularly the southern rock lobster and abalone industries, worth more than AU$ 160 million annu- ally (Johnson et al. 2005, Ling et al. 2009a, ABARES 2011).
Temperature, nutrients and light generally limit kelp to mid-latitudes of between 40° and 60° in the Northern and Southern hemispheres (Steneck et al. 2002), although northern range limits of Ecklonia radiata in Australia extend to ~28°S (Guiler 1960, Kirkman & Kendrick 1997). Changes to temperature and nitrates can impair kelp photosynthetic, respira- tory and cellular function in their expected ranges (Staehr & Wernberg 2009, Wernberg et al. 2010), leading to reduced growth and survivorship (Dayton et al. 1999, Graham et al. 2007) and ultimately range contractions or local extirpations (van Tussenbroek 1989, Edyvane 2003, Connell et al. 2008). Moreover, increases in ocean temperature associated with cli- mate change have been linked to reduced resilience in kelp to anthropogenic and natural perturbations (Wernberg et al. 2010). Understanding the combined effects of temperature and nitrates on E. radiata will provide valuable information towards understanding climate change impacts and the future viability of E. radiata populations in southeastern Australia. In par- ticular, an understanding of the impacts of climate change on the microscopic stages (gametophyte and juvenile sporophytes) of the life cycle is likely to be crucial (Hurd et al. 2009, Roleda et al. 2012), but to our knowledge, combined impacts of changes in temperature and nutrients on microscopic stages have not been examined in E. radiata.
Given the observed changes to the EAC and their influence on local-scale temperature and nitrates, there are likely to be important implications for the reproduction, development and growth of kelp. Our overall aim was to examine how the predicted scena - rio of a stronger EAC may influence the reproduction and development of the dominant habitat-forming kelp in temperate Australia, Ecklonia radiata. We addressed this aim by combining large-scale field surveys of E. radiata morphology and reproduction in 2 latitudinal regions in southeastern Australia (Tas- mania and NSW) with laboratory culturing experi- ments to determine the effects of temperature and nitrates on different microscopic life-cycle stages. Specifically, we determined (1) the daily SST and wave power at sites in Tasmania and NSW, (2) differ- ences in the morphology and the amount of repro- ductive tissue of E. radiata sporophytes among sites in Tasmania and NSW, (3) whether variation in the amount of reproductive tissue was explained by vari- ation in thallus morphology, and (4) how increasing
118
temperatures and lower nitrates affected the growth and development of gametophytes and juvenile sporophytes of E. radiata.
MATERIALS AND METHODS
Study species
Ecklonia radiata (C. Agardh) J. Agardh is a perennial, subtidal, habitat-forming kelp (order Lamina riales; Womersley 1967) with a wide lat- itudinal and depth range (Womersley 1981), making it the most important habitat-forming species in southeastern Australia. It has a typi- cal laminarian life cycle consisting of micro- scopic male and female gametophytes and a macroscopic sporophyte stage. Sporophytes grow up to 2 m in length and consist of a stipe arising from a single conical holdfast. The cen- tral lamina (primary blade) arises from the stipe and basal meristem, with laterals (secondary blades) developing in distichous pairs along the central lamina (Womersley 1987). The lamina can be eroded by forces from wave action and grazing (Mann 1973, Kirkman 1981). Adult sporophytes develop sori along the central lam- ina and laterals typically at approximately 8 to 12 mo post fertilisation (Novaczek 1980, Kirk- man 1981). Sori emerge from the distal (top) end of the thallus and grow down the central lamina and into the laterals towards their distal ends. Zoospores are released from sori, and after settlement, zoospores germinate into 1-celled male and female gametophytes that become fertile through gametogenesis. Game- tophytes are dioecious and filamentous, ex - hibiting sexual dimorphism, and females (up to 400 µm in length) are larger than males (up to 100 µm in length) and less branched. Gameto- phytes are oogamous, and male gametophytes produce motile antherozoids that fertilise non- motile eggs which have been liberated from elongated terminal cells, the oogonia (Womersley 1981), that then metamorphose into juvenile sporo- phytes (Womersley 1981). E. radiata is known to sur- vive up to 10 yr in New Zealand (Novaczek 1981).
Study sites and sampling
We determined SST and wave power and sampled adult sporophytes, which we defined as reproductive
thalli or thalli of equivalent size and morphology to reproductive thalli, from 3 sites within each of 2 regions spanning ~10° of latitude (NSW and Tasma- nia, Fig. 1). Adult sporophytes were collected hap- hazardly from largely monospecific Ecklonia radiata forests between 7 and 10 m depth from sites exposed to year-round prevailing south to southeasterly swells (Guiler 1960, Wright 1976). As much as possi- ble, we endeavoured to standardise large-scale envi- ronmental factors that are known to influence kelp
119
Fig. 1. (A) Temperature profiles of the 6 study sites from December 2008 to September 2010; Tasmania (grey dots): Fortescue Bay (FB), Bicheno (BO), Eddystone Point (EP); New South Wales (black dots): Bass Point (BP), Bare Island (BI), Long Bay (LB). (B) Map showing the 6 sites in southeastern Australia where Ecklonia radiata were sampled for sea surface temperature, morphology and reproduc- tion, with insets showing average wave power (1997 to 2008);
pixels indicate average W m–1 of wave crest
Mar Ecol Prog Ser 483: 117–131, 2013
morphology (see Fowler-Walker et al. 2005, 2006), as local-scale drivers of morphology such as hydrology and sedimentation are highly complex to determine and measure.
Variation in SST and wave power
SST data were collected for the 6 sites using day- time data sourced from Bands 31 and 32 of the mod- erate resolution imaging spectroradiometer aboard the Aqua satellite, which is an effective proxy for in situ SSTs for depths of up to 10 m (Smale & Wernberg 2009). Image data are administered by Ocean Color Web, Ocean Biology Processing Group (OBPG) at NASA’s Goddard Space and Flight Centre, Green- belt, Maryland, USA (Feldman & McClain 2010) and were downloaded from the OBPG Web site (http:// oceancolor.gsfc.nasa.gov/). Temporal and spatial res- olutions are daily at 4 × 4 km. Using Matlab R2009b, data were extracted from the images (Feldman & McClain 2010) using a 5 × 5 pixel (20 × 20 km) win- dow around each of the sampling sites, and data were interpolated by averaging temperature values within this region. Daily SST data were extracted for the previous 12 mo mean at each site for each sam- pling period. This 12 mo time period was chosen based on time to adulthood in Ecklonia radiata (Novaczek 1980, Kirkman 1981).
Wave power (W m−1 crest), estimated by the regio - nal implementation of the WAM model (Greenslade 2001), provides an estimate of the time-averaged wave power that occurs on the Australian shelf. The WAM model integrates the basic transport equation describing the evolution of a 2-dimensional ocean wave spectrum, including dissipation due to white- capping (Hasselmann 1988). Linear wave theory and water depth were used to calculate wave power from the WAM-derived significant wave height and period (Hughes & Heap 2010). The estimated time- average wave power was obtained by averaging the power record over the available time series. Data were extracted from the model output to describe the time-average wave power at each of the sample sites.
Variation in morphology and reproduction
Morphology and reproduction of adult Ecklonia radiata sporophytes were quantified at each site on 3 occasions (May, July and September 2010). Each time, 23 to 24 Stage 3 E. radiata sporophytes (see Kirkman 1981, 1984) were collected (no collection
was done in July from Bass Point due to bad weather). Thalli were returned to the lab, where 7 morphological traits (May and July only) and 1 repro- ductive trait were measured (Fig. 2). The 7 morpho- logical traits (thallus length, lamina length, lamina width, lateral length, lateral width, stipe length and stipe thickness) reflect the overall size and shape of E. radiata thalli and are a subset of traits measured previously by others (see Wernberg et al. 2003, Fowler-Walker et al. 2005, 2006, Wernberg & Thom- sen 2005). The total amount of reproductive tissue down the central lamina of a thallus was measured as the distance from the distal end where sori emerge to where sori terminate (Fig. 2). We focused on this measure of reproductive tissue, as the majority of reproductive tissue occurs down the central lamina (~90%) compared to the laterals; sori coverage on laterals was patchy, and sori-bearing laterals always occurred within the margins of central lamina sori. While this measure does not reflect all aspects of reproduction (Patten & Yarish 1993), it is a meaning- ful relative measure that allows quantitative among- site com parisons. Moreover, there is a link between sorus area and spore release density (Mohring et al. 2013). In Tasmania, E. radiata reproduction output peaks between May and July (Sanderson 1990); how ever, this is not known for NSW. Thus, our sam- pling coincided with the time when reproduction was
120
Fig. 2. Ecklonia radiata. Seven morphological measure- ments taken on adult sporophyte: (a) total thallus length; (b) stipe length; (c) central lamina length; (d) central lamina width; (e) stipe diameter; (f) lateral length; (g) lateral width.
Sori extent down the central lamina is also shown (h)
Mabin et al.: Kelp responses to changing ocean conditions
high and is a meaningful test of the link between variation in morphology and reproduction. In Sep- tember, we only measured lamina length and stipe length, focusing on those morphometrics that were good predictors of reproduction.
Testing for effects of temperature and nitrates on growth and development
Microscopic stages of Ecklonia radiata were grown from zoospores liberated from reproductive tissue (sorus) from 10 reproductive sporophytes collected at Bicheno in July 2010. To liberate zoospores, sori were placed into a cool, dry and dark place for approxi- mately 1 h and then briefly sterilised in an antiseptic solution (1 ml Betadine l−1 distilled water) for ~1 min so as not to harm the spores whilst reducing contami- nation by associated epiflora (Wright et al. 2004). Ex- cessive mucilage was wiped away with a lint-free cloth. Sori were then submerged into 500 ml of 0.2 µm filtered, autoclaved seawater and placed under con- trolled light conditions. Zoospores were released within 7 h. Then, 100 µl of zoospore solution (~2.8 × 105 zoospores ml−1, pooled from all 10 plants) was pipetted into cylindrical, autoclaved plastic culture jars (43 mm diameter, 45 mm depth) containing seawater media (see below) and 2 coverslips as the settlement substratum, to be used for destructive sampling of microscopic parameters at 2 times, post- settlement (gametophytes) and post-fertilisation (sporo phytes). Jars (n = 3 replicates of each tempera- ture × nitrate concentration combination) were placed into an aluminium temperature-gradient plate (730 × 370 mm) forged with 36 wells (6 × 6 grid, 44 mm di- ameter, 45 mm depth). Lighting consisted of 4 parallel 40 W Sylvania standard cool white globes (Model F40W/133-RS) set to a 16 h light:8 h dark cycle, the optimum lighting re gime for E. radiata development (Novaczek 1984a). Light flux density ranged from ~17 to ~42 µmol photons m−2 s−1. The growth and devel- opment of E. radiata gametophytes do not differ within this light range (Novaczek 1984a,b).
A temperature gradient across the aluminium plate was maintained by water baths at either end (12 and 26°C), resulting in 6 experimental temperatures: 13.5 ± 0.1°C, 15 ± 0.1°C, 16.5 ± 0.1°C, 19 ± 0.1°C, 22 ± 0.1°C, and 25.5 ± 0.1°C. High temperatures reflected the predicted average SST increase of 2 to 3°C off the Tasmanian coast by the year 2100, based on oceano- graphic CO2 models (Ridgway 2007, Ridgway & Hill 2009). The second highest temperature (22.0°C) ful- filled this objective, while the maximum temperature
(25.5°C) tested for the effects of an extreme increase in temperature on the microscopic stages of Tasma - nian Ecklonia radiata.
Two nitrate concentration treatments of 0.5 µmol l−1
and 3.0 µmol l−1 were created using f/2 media (re - viewed by Andersen 2005), where sodium nitrate (NaNO3) was adjusted accordingly to achieve the desired nitrate concentrations. Nitrate levels were set to reflect normal and depleted dissolved nitrate lev- els off Maria Island (Rochford 1984). Sodium dihy- drogen phosphate di hydrate (NaH2PO4 · 2H2O) was also adjusted to retain the N:P ratio of the f/2 media. Silicates were omitted from the recipe to reduce the likelihood of diatom contamination, which elimi- nated the need to use germanium dioxide (GeO2). GeO2 is used as a diatom suppressant in Laminaria culturing due to its effect of arresting diatom cell division (Lewin 1966, Shea & Chopin 2007); however, GeO2 may have inhibitory effects on kelp gameto- phytes (Shea & Chopin 2007). Media was changed weekly.
One coverslip was removed from each culture jar at 16 d (post-settlement, for male and female gameto- phytes) and the second at 32 d (post-fertilisation, for sporophytes) to determine growth and development. Once removed, coverslips were inverted onto a mi- croscope slide, and up to 100 photographs were taken haphazardly from each coverslip at a magnification of between 200× and 400×. Images were randomly se- lected until 5 male and 5 female gametophytes were measured per coverslip (n = 15 per sex, per crossed nitrate × temperature treatment). Two measurements were made on each gametophyte; 2-dimensional sur- face area and the number of cells of the gametophyte. Surface area was measured in ImageJ 1.43 using the trace tool with the scale calibrated to 0.1 mm using a digital image of a calibra ting slide (Wild Heerbrugg, Model 310345). Cells were counted manually from images. Only surface area was measured on juvenile sporophytes (3 per coverslip, n = 9 per crossed treat- ment) using the same methods as those used for ga- metophytes. Surface area measures growth (Shea & Chopin 2007), while cell count measures de velop - ment via cell division rates (Novaczek 1980) under the different treatments.
Statistical analyses
Multivariate differences in Ecklonia radiata mor- phology were examined using nested permutational multivariate ANOVA (PERMANOVA; Anderson 2001) with the factors Region (random), Site (ran-
121
Mar Ecol Prog Ser 483: 117–131, 2013
dom) nested within Region and Time (fixed). We used untransformed Gower dissimilarities (Gower 1967), as we had variables measured at different scales and ran 9999 permutations to calculate F-sta- tistics. Permutational tests of multivariate dispersion (PERMDISP; Anderson et al. 2008) were done to assess the effect of the dispersion of observations about their centroid on the centroid’s position in mul- tivariate space. Finally, pseudo-variance components were then calculated for each of the spatial levels each month (region, sites and within sites) (e.g. Benedetti-Cecchi 2001). To visualise multivariate dif- ferences in E. radiata morphology, non-metric multi- dimensional scaling (nMDS) was done using Gower’s dissimilarities. Ordinations were based on average centroids for sites each month to limit the number of observation points on plots. All multivariate analyses were run using PERMANOVA+ for PRIMER. Differ- ences in individual morphological traits were deter- mined using 3-factor nested ANOVA with the factors (all random) Region (Tasmania vs. NSW), Site (nes - ted within Region) and Time (May and July 2010).
Differences in reproduction (extent of sori down the central lamina) between regions and among sites within regions at each time (May, July and Septem- ber) were determined using nested analysis of co - variance (ANCOVA). The effect of time was not included in this analysis. Total length of the central lamina was used as a covariate to take into account effects of lamina size on reproductive tissue. To meet ANCO VA assumptions, data were logn(x + 1) trans- formed and homogeneity of slopes was examined. Despite the transformation, this assumption was vio- lated in July and September (see ‘Results’).
To test for relationships between the extent of sori down the central lamina and particular morphologi- cal traits of thalli, multiple regressions were per- formed separately for May and June, with the extent of sori down the central lamina as the dependent variable and all morpho- logical measures, ex cluding total thallus length, as predictor variables. Data were logn(x + 1) transformed.
Differences in the surface area and cell count of gametophytes were ana - lysed separately with 4-factor nested ANOVA with the factors Tempera- ture (fixed), Nitrate concentration (fixed), Sex (fixed) and Jar (random and nested within the Temperature × Nitrate interaction). Both these de - pendent factors were logn(x + 1) transformed. Differences in the sur-
face area of sporophytes (logn transformed) was analysed with 3-factor nested ANOVA with the fac- tors Temperature (fixed), Nitrate (fixed) and Jar (ran- dom and nested within the Temperature × Nitrate interaction).
RESULTS
Variation in SST and wave power: Tasmania vs. NSW
Average SST in NSW and Tasmania differed by approximately 5°C. NSW sites showed a minimum of 13.7°C, a maximum of 24.4°C and a mean of 19.3°C (±0.07°C). Tasmanian sites ranged between 8.4 and 19.9°C with a mean of 14.5°C (±0.07°C). Strong sea- sonal patterns occurred for sites in both regions, with maximum temperatures in February to March and minimum temperatures in June to August (Fig. 1A). Average wave power from 1997 to 2008 varied among sites and was highest at Bass Point (NSW) (9285.03 W m−1 crest) and lowest at Fortescue Bay (3350.57 W m−1 crest). Generally, average wave power was greater at NSW sites (7549.543 ± 867.74 W m−1, mean ± SE) compared to Tasmanian sites (5326.64 ± 1120.84 W m−1, mean ± SE), although the magni- tude of difference was small (Fig. 1B).
Differences in morphology between regions
PERMANOVA indicated significant variation in kelp morphology at the scales of Region, Site within Region and a significant Site within Region × Time interaction (Table 1). Pseudo-variance components highlighted large spatial variation at the scales of Region and Site within Region and, in particular, at
122
Source df MS Pseudo-F p % of total (perm) variance
Region 1 10313 4.829 0.032 23.7 Site (Region) 4 2322.1 15.758 <0.001 17.3 Time 1 2361.1 3.967 0.328 5.1 Region × Time 1 595.14 1.468 0.297 1.1 Site (Region × Time) 3 405.3 2.751 0.006 3.6 Residual 253 147.36 49.2
Table 1. Ecklonia radiata. Results of nested PERMANOVA testing the effects of Region (Tasmania vs. New South Wales), Sites within Region and Time (May and July 2010) on morphology. There were 3 sites per region except for New South Wales in July, when 1 site could not be accessed. Significant
(p < 0.05) factors in bold
Mabin et al.: Kelp responses to changing ocean conditions
the smallest spatial scale (among individuals within Sites), which accounted for >50% of the variation (Table 1). Visual inspection of nMDS plots reflected the results from the PERMANOVA; although there was some evidence of grouping of sites within regions each month, there was still large separation among sites within regions (Fig. 3).
Most individual morphological traits showed signi - ficant variation among sites within regions (Table 2, Fig. 4). Two traits that reflect overall thallus size (total length and lateral length) were significantly larger in Tasmania compared to NSW, while 2 other traits (lamina width and lamina length) varied across re - gions and time (significant Region × Time interac- tion, Table 2). Lateral width was also highly variable [significant Site (Region) × Time interaction], but stipe diameter did not vary with any factors (Table 2).
Differences in reproduction between regions
In May, adult sporophytes from Tasmania had sig- nificantly more reproductive tissue on the central lamina than plants from NSW (region, F1,4 = 6.032, p = 0.028; site, F4,136 = 0.499, p = 0.736; lamina length, F1,136 = 49.484, p < 0.001; Fig. 5). By contrast, in both
July and September, there were no differences be - tween regions but significant differences among sites within regions (July: region, F1,3 = 0.626, p = 0.488; site, F3,115 = 5.101, p = 0.002; lamina length, F1,115 =
123
Source df MS F p
Total length Region 1 31292.6 23.431 0.008 Time 1 9481.60 8.761 0.207 Site (Region) 4 1335.50 5.365 <0.001 Region × Time 1 1082.30 4.437 0.126 Site (Region) × Time 3 243.90 0.980 0.403 Residual 251 248.90
Stipe length (log10) Region 1 4.028 5.379 0.081 Time 1 0.450 0.810 0.534 Site (Region) 4 0.749 10.611 <0.001 Region × Time 1 0.556 5.836 0.095 Site (Region) × Time 3 0.095 1.348 0.259 Residual 251 0.071
Lamina length (log10) Region 1 1.152 1.785 0.253 Time 1 0.545 1.690 0.417 Site (Region) 4 0.645 40.476 <0.001 Region × Time 1 0.323 55.731 0.005 Site (Region) × Time 3 0.006 0.363 0.780 Residual 251 0.016
Lamina width (log10) Region 1 0.175 0.422 0.558 Time 1 0.134 0.992 0.604 Site (Region) 4 0.415 45.082 <0.001 Region × Time 1 0.135 345.846 <0.001 Site (Region) × Time 3 0.000 0.042 0.988 Residual 251 0.009
Stipe diameter Region 1 0.029 0.232 0.655 Time 1 0.315 3.629 0.307 Site (Region) 4 0.126 2.108 0.080 Region × Time 1 0.087 0.584 0.500 Site (Region) × Time 3 0.149 2.487 0.061 Residual 251 0.060
Lateral length Region 1 9411.50 66.512 0.001 Time 1 260.200 2.694 0.348 Site (Region) 4 141.500 1.762 0.137 Region × Time 1 96.600 0.477 0.539 Site (Region) × Time 3 202.600 2.522 0.058 Residual 251 80.300
Lateral width (log10) Region 1 1.099 3.310 0.143 Time 1 0.266 10.161 0.194 Site (Region) 4 0.332 22.045 <0.001 Region × Time 1 0.026 0.386 0.579 Site (Region) × Time 3 0.068 4.508 <0.001 Residual 251 0.015
Table 2. Ecklonia radiata. ANOVA testing the effects of Region (Tasmania vs. New South Wales), Site within Region and Time (May and July 2010) on 7 morphological traits of
morphology. Significant (p < 0.05) factors in bold
BI
LB
BO
EP
Fig. 3. Ecklonia radiata. Non-parametric multidimensional scaling analyses showing morphological variation at sites in (A) May and (B) July. Site abbreviations as in Fig. 1. Filled
symbols: Tasmania; open symbols: New South Wales
Mar Ecol Prog Ser 483: 117–131, 2013124
Total thallus length S
100
80
60
40
20
0
12
10
8
6
4
2
0
50
40
30
20
10
0
50
40
30
20
10
0
10
8
6
4
2
0
2
1
1
0
60
40
20
0 BO EP BP BI LB FB BO EP BP BI LB FB BO EP BP BI LB
FB BO EP BP BI LB
FB BO EP BP BI LB
FB BO EP BP BI LB FB BO EP BP BI LB
Site
Site
Lamina length
Lamina width
Fig. 4. Ecklonia radiata. Mean (±1 SE) of 7 morphological traits among sites in May (light grey bars) and July (dark grey bars). Tasmanian sites: Fortescue Bay (FB), Bicheno (BO), Eddystone Point (EP); New South Wales sites: Bass Point (BP), Bare Island (BI), Long Bay (LB). *Site not sampled
LBBIBP
Site Site Site
Fig. 5. Ecklonia radi- ata. Amount of repro- ductive tissue down the central lamina of thalli at sites in May, July and September. Site abbreviations as in Fig. 4. *Site not sampled. Bars show
mean ±1 SE
Mabin et al.: Kelp responses to changing ocean conditions
21.924, p < 0.001; September: region, F1,4 = 0.872, p = 0.403; site, F4,136 = 4.084, p = 0.004; la mina length, F1,136
= 31.036, p < 0.001). In June and Sep- tember, there were also violations of the ANCOVA homogeneity of slope assumption, highlighting the large site-to-site variation in reproduction but limiting the interpretation for these months (see ‘Discussion’). For all sites, reproduction declined over the duration of the study (May to Sep- tember) with the exception of Bare Island (NSW) from May to July. Reproduction then de clined at this site from July to September (Fig. 5).
Do morphological traits predict reproduction?
In May, approximately 73% of the variation in the extent of reproduction tissue down the central lamina was explained by kelp morphology (R2 = 0.727, F6,136
= 60.469, p < 0.001), while in July, only 37% of the variation in the extent of reproduction tissue down the central lamina was explained this way (R2 = 0.374, F6,113 = 11.234, p < 0.001). Of all predictor vari- ables, lamina length and stipe diameter were the only two with significant and positive linear relation- ships with the amount of reproduction. Examination of the partial residual plots for these 2 factors against reproduction showed that only lamina length was a strong and significant predictor of the extent of reproduction tissue down the central lamina (May: R2
= 0.667, p < 0.001; July: R2 = 0.324, p < 0.001), while stipe diameter, although significant, was weakly cor- related with reproduction (May: R2 = 0.043, p = 0.015; July: R2 = 0.054, p = 0.012).
Effects of temperature and nitrates on growth and development
The surface area of gametophytes 16 d after spore release differed significantly between sex and tem- perature (Table 3). Males had a larger surface area than females, while at temperatures less than 15°C and greater than 22°C, gametophytes were signifi- cantly smaller than at temperatures between 15 and 22°C (Fig. 6). Noticeably, gametophytes grown at 25.5°C were significantly smaller in surface area than those grown at all other temperatures. There was no effect of nitrates on growth and no sig nificant Tem-
125
Source SS df MS F p
Temperature 7.344 5 1.469 32.219 <0.001 Nitrate 0.117 1 0.117 2.574 0.122 Sex 0.472 1 0.472 15.609 0.001 Temperature × Nitrate 0.185 5 0.037 0.810 0.554 Temperature × Sex 0.108 5 0.022 0.713 0.620 Nitrate × Sex 0.047 1 0.047 1.561 0.224 Temperature × Nitrate × Sex 0.173 5 0.035 1.141 0.366 Jar (Temperature × Nitrate) 1.094 24 0.046 1.506 0.161 Sex × Jar (Temperature × Nitrate) 0.726 24 0.030 1.539 0.054 Residual 5.664 288 0.020
Table 3. Ecklonia radiata. ANOVA testing the effects of temperature, nitrates and sex on gametophyte surface area with jar replicates included as a nested term. Surface area data were log10 transformed prior to analysis to meet ANOVA homogeneity of variance assumptions. Significant (p < 0.05) factors
in bold
b
c
d
d
c,d
a
S ur
fa ce
a re
a (µ
m 2 )
C el
nt
2000
1500
1000
500
0
30
20
10
0
Fig. 6. Ecklonia radiata. (A) Variation in surface area of male and female gametophytes across 6 temperatures; (B) varia- tion in cell count of male and female gametophytes across 6 temperatures. Bars show mean ±1 SE. For (A), means shar- ing letters are not significantly (p > 0.05) different from one
another
Mar Ecol Prog Ser 483: 117–131, 2013
perature × Nitrate in ter action. In con- trast to surface area, the number of cells per individual was highly vari- able. There was a significant Temper- ature × Nitrate × Sex interaction (Table 4), indicating that within each Temperature × Nitrate × Sex combi- nation cell count was variable and the main effects above this term for cell count cannot be further interpreted. There was also significant random variation among jars for each treat- ment combination [significant Sex × Jar (Temperature × Nitrate)] (Table 4).
For sporophyte surface area, there were significant differences among temperatures but no significant effect of nitrate and no significant Tempera- ture × Nitrate interaction. Most not - ably, no sporophytes developed in the 25.5°C treatment. Sporophytes at 16.5°C had a significantly larger sur- face area than those of the 2 lower temperatures, while there were no sig- nificant differences in surface area for sporophytes grown in the 16.5°C, 19°C and 22°C treatments (Table 5; Fig. 7).
DISCUSSION
Globally, the world’s oceans are becoming warmer and major ocean currents are shifting, resulting in demographic and distributional shifts in important marine species (Parmesan 2006, Poloczanska et al. 2007). Stronger EAC penetration into southeastern Australian coastal waters is already causing shifts in species distributions (Johnson et al. 2005, Ling et al. 2009a,b, Johnson et al. 2011, McLeod et al. 2012a,b), with SST in the region predicted to increase by up to 3.0°C by the year 2100 (Rochford 1984, Ridgway 2007). Be cause increased temperatures and reduced nitrates represent forcing factors for kelp, under- standing how they impact the reproduction and development of Ecklonia radiata, the most abundant and widely distributed canopy-forming kelp in Aus- tralia, is important. Our findings show clear effects of increasing temperature (but not reduced nitrates) on the development and growth of microscopic life- cycle stages and indicate that because reproduction is linked to morphology, any temperature- or nitrate- driven changes to morphology may also have impli- cations for reproduction.
126
Source SS df MS F p
Temperature 5.443 5 1.089 52.181 <0.001 Nitrate 0.121 1 0.121 5.788 0.024 Temperature × Nitrate 0.303 5 0.061 2.903 0.035 Sex 28.338 1 28.338 1341.175 <0.001 Nitrate × Sex 0.008 1 0.008 0.370 0.549 Temperature × Sex 1.257 5 0.251 11.903 <0.001 Temperature × Nitrate × Sex 0.319 5 0.064 3.018 0.030 Jar (Temperature × Nitrate) 0.501 24 0.021 0.987 0.512 Sex × Jar (Temperature × Nitrate) 0.507 24 0.021 1.731 0.020 Residual 3.516 288 0.012
Table 4. Ecklonia radiata. ANOVA testing the effects of temperature, nitrates and sex on gametophyte cell count with jar replicates included as a nested term. Cell count data were log10 transformed prior to analysis to meet ANOVA homogeneity of variance assumptions. Significant (p < 0.05) factors in bold
Source SS df MS F p
Temperature 187.269 5 37.454 83.501 <0.001 Nitrate 1.003 1 1.003 2.236 0.148 Temperature × Nitrate 5.789 5 1.158 2.581 0.053 Jar (Temperature × Nitrate) 10.765 24 0.449 0.493 0.973 Residual 65.449 72 0.909
Table 5. Ecklonia radiata. ANOVA testing the effects of temperature and nitrates on sporophyte surface area with jar replicates included as a nested term. Sporophyte surface area data had heterogeneous variances (Cochran’s C-test: p < 0.05) and were log10 transformed prior to analysis to meet ANOVA homogeneity of variance assumptions. Significant (p < 0.05) factors in bold
13.5 15.0 16.5 19.0 22.0 25.5
a a
800
600
400
200
0
Fig. 7. Ecklonia radiata. Variation in the surface area of microscopic sporophytes in 6 treatments of temperature. Bars show mean ±1 SE. Means sharing letters are not signi-
ficantly (p > 0.05) different from one another
Mabin et al.: Kelp responses to changing ocean conditions
Morphology
The morphology of Ecklonia radiata was highly variable within and among 6 eastern Australian sites, in line with previous broad-scale observations of E. radiata morphology among NSW, South Australia and Western Australia (WA) (Wernberg et al. 2003, Fowler-Walker et al. 2005) and along a 6° latitudinal gradient on the WA coast (Wernberg et al. 2010). Our study incorporated a strong latitudinal aspect (~8.5°), and some differences in traits occurred between NSW and Tasmanian sites. This suggests that al - though there is high local-scale variation in E. radiata morphology, there are also regional-scale differ- ences, as suggested by previous studies (Chesson 1996, Fowler-Walker et al. 2005). Tasmanian popula- tions of E. radiata typically had a larger morphology compared to NSW populations, especially for thallus length and lateral length. Nonetheless, the large variation in E. radiata morphology within and among sites was notable; a striking example of this occurred between the 2 closest sites in NSW, Bare Island and Long Bay, for thallus length and lamina length. This local variation makes interpreting the overall causes of local-scale morphological patterns difficult but highlights the importance of small-scale processes such as wave exposure, water depth (Fowler-Walker et al. 2005) and nitrates that may be associated with localised anthropogenic input. Currently, there ap - pears to be slightly greater wave power in NSW than eastern Tasmania. Given that wave height and speed are forecast to increase globally (Young et al. 2011), it is likely that wave activity will continue to modify the morphology of wave-exposed E. radiata.
Temperature and nitrates, among other factors, are recognised as key drivers for large-scale growth and, hence, morphological differences in adult lamina - rians (Dayton 1985, Fowler-Walker et al. 2005, Gra- ham et al. 2007), with age often being a weaker pre- dictor of growth and morphology (Chapman 1986). For example, in WA, Ecklonia radiata growth rates are reduced when water temperatures exceed 20°C (Hatcher et al. 1987) and adult E. cava transplanted from warmer to cooler water increased in size rela- tive to their non-transplanted cohorts, highlighting a plastic response in adults to temperature (Serisawa et al. 2002). The morphological differences between Tasmania and NSW (annual mean SSTs: Tasmania ~14.0°C; NSW ~19.5°C) suggest both morphological and growth responses to temperature variation. However, SST may not be the only large-scale con- tributing forcing factor to the patterns observed. On regional scales, nitrates may also influence growth
and morphology, and in the field, they are negatively correlated with temperature. In situ nitrate addition experiments with Macrocystis pyri fera demonstrated increased survivorship, growth rates and larger phe- notypic plasticity in response to high nitrate levels compared to M. pyrifera in low nitrate (and higher temperature) conditions associated with El Niño events in California (Dayton 1985, Dean & Jacobsen 1986).
Furthermore, the extent to which the morphologi- cal differences reflect genetic vs. environmental influences is not clear, as the heritability of these traits is not known. There are genetic differences between Ecklonia radiata from Tasmania and NSW (Dolman & Coleman 2008), but phenotypic plasticity for Ecklonia morphology has been demonstrated in transplanted individuals (Serisawa et al. 2002, Fowler-Walker et al. 2006). Regardless of the degree to which morphology is plastic or heritable, because E. radiata generally have a smaller size in the warmer waters of NSW than in Tasmania, increased temperatures from EAC strengthening may contrib - ute to a reduction in adult E. radiata size in south- eastern Australia. Under such a scenario, there are implications for kelp forest architecture and poten- tially how E. radiata engineers the surrounding envi- ronment (Wernberg & Thomsen 2005) and impacts associated species.
Reproduction
Although quantifying the amount of reproductive tissue per thallus by measuring the extent of repro- ductive tissue down the central lamina, from the dis- tal end of the thallus, is only one method for quantify- ing reproduction (Patten & Yarish 1993), it allowed for a rapid and relative comparison of reproduction (as a function of lamina length) over biogeographical scales. Laminarians generally exhibit seasonal varia- tion in reproductive output which can vary with loca- tion and depth due to changes in temperature and light (Bolton & Levitt 1985, Joska & Bolton 1987). Reproductive seasonality in Ecklonia radiata has been previously recorded in Tasmania (Sanderson 1990) and WA (Mohring et al. 2013) and is likely to occur in NSW. The differences in E. radiata morphol- ogy are important, as central lamina length was a strong predictor of reproduction in May when repro- duction was highest. Lamina length was still a strong predictor of reproduction in July and September when reproduction declined and when the ANCOVA assumption of homogeneity of slopes was violated.
127
Mar Ecol Prog Ser 483: 117–131, 2013
For July and September, heterogeneous slopes are interpreted as the relationship be tween reproduction and the covariate (lamina length) differing between regions and/or among sites within regions, highligh - ting the high variability for reproduction in these months. This variability may be an effect of sampling outside the main reproductive window of E. radiata at these locations (Mohring et al. 2013). Positive size- fecundity relationships often occur in sessile organ- isms including plants (Aarssen & Clauss 1992) and laminarians (Reed et al. 2004, Underwood et al. 2000). The regional differences in reproduction may also reflect a response to higher mean annual tem- peratures (i.e. NSW) as observed for other kelp (e. g. Laminaria longicruris, Patten & Yarish 1993), a re - sponse to lower nitrates as observed in Macrocystis pyrifera (Reed et al. 2004) or a response to a combi- nation of both factors (Dayton 1985, Dean & Jacobsen 1986). Because lamina length covaries with repro- duction, smaller size in E. radiata under higher tem- peratures suggests reduced reproductive capacity.
Effects of temperature and nitrates on Ecklonia radiata development
Temperature also had strong effects on the de - velopment of microscopic stages of Ecklonia radiata. Both the gametophyte stage (measured at 16 d post- settlement) and sporophyte stage (measured at 32 d post-settlement) experienced optimum growth be - tween temperatures of 16.5 and 22.0°C. In subopti- mal temperatures (above 22.0°C), gametophytes grew the least, and no sporophytes had emerged by Day 32, highlighting a potential temperature thresh- old for E. radiata development. The 25.5°C tempera- ture treatment reduced but did not stop gametophyte growth. The absence of sporophytes at this tempera- ture does not necessarily deem them incapable of growth at this temperature but suggests that no suc- cessful fertilization occurred. The mechanism by which this may happen is not known but may result from arrested development of male or female ga - metes due to gametes being infertile or not function- ing normally (e.g. antherozoids being unable to swim or respond to pheromones) or post-fertilisation mor- tality (Muñoz et al. 2004). This may also reflect a tem- perature-mediated seed bank strategy in E. radiata, with gametogenesis occurring within a narrower temperature range than gametophyte growth and survival (tom Dieck 1993). Furthermore, the optimum temperature range may influence the timing of ferti - li zation, as gametogenesis may be a function of tem-
perature. We were unable to determine the precise moment of temperature-dependent gametogenesis from our photographs, but this is an area for future research.
As nitrates did not have a significant impact on gametophytes and sporophyte growth, temperature- driven latitudinal effects appear more important and may determine upper survival temperature thresh- olds for microscopic Ecklonia radiata in Tasmania. An upper survival temperature of 28°C was found previously for Tasmanian E. radiata gametophytes (tom Dieck 1993). In New Zealand, E. radiata game- tophytes sourced from cooler waters exhibited lower maximum temperature thresholds and followed the same pattern of reduced growth and no reproduction when exposed to such temperature extremes (Nova - czek 1984a), suggesting localised adaptation to local temperature. Our cultured spores were sourced from Stage 3 adults in Bicheno, Tasmania, at a depth of 10 m (latitude ~42°S), where temperatures fluctuate annually around a mean of ~15°C with a summer maxima of ~19.0°C. Upper survival temperature thresholds may be different for gametophytes and juvenile sporophytes (Novaczek 1984a, tom Dieck 1993) in NSW, where the average annual tempe - rature is ~20°C. Examining the tolerance of game - tophytes and juvenile sporophytes sourced from multiple locations across latitudes, to the same exper- imental temperatures and nitrate concentrations, could determine whether similar tolerances occur or whether juveniles from warm-water locations are adapted to those conditions. In contrast to other experiments where optimum temperatures were found for kelp gametophyte cell count (e.g. E. max- ima, Bolton & Levitt 1985), there was high variability in gametophyte cell count across treatments, sug- gesting variable levels of cell division and growth. The number of cells in male gametophytes peaked at 19°C and presumably allows an excess of antheridia production and a higher probability of fertilisation success. The absence of a difference in the surface area of gametophytes or sporophytes between high and low nitrates concurs with results for high (9 µmol l−1 NO3) and low (~2 µmol l−1 NO3) nutrient effects on the surface area Macrocystis pyrifera sporophytes in California under different light conditions (see Kin- lan et al. 2003) but contrasts to findings for M. pyri - fera in Chile, where females grew larger at high nitrate concentrations (although nitrate concentra- tions were not stated, Muñoz et al. 2004).
Generally, Laminariales gametophytes have a higher temperature tolerance than macroscopic adults (sporophytes) (tom Dieck 1993). If experimen-
128
Mabin et al.: Kelp responses to changing ocean conditions
tal results for microscopic stages of Ecklonia radiata are translated into a field scenario, the temperature tolerance of macroscopic adults may be a limiting factor for E. radiata distribution. However, the find- ing of an upper temperature threshold for sporophyte development suggests temperature effects on the microscopic stages may be critical.
CONCLUSION
Climate-induced changes to ocean temperatures and currents which affect important habitat-forming species, such as Ecklonia radiata, will have broad impacts on coastal food webs. Our study highlights some of the likely consequences for E. radiata mor- phology, reproduction and development given future climate scenarios. Such information is im - portant for predicting impacts of EAC strenghtening on southeast Australian kelp forests, and can con- tribute to reducing uncertainty for stakeholders with a vested interest in kelp ecosystems, whilst con- tributing information towards human adaption to climate change.
Acknowledgements. We thank the following people for their assistance in undertaking this research: M. Filleul, D. Harri- son, K. Leonard, E. Flukes, D. Bryant, I. Jermyn, C. Bolch and C. Sanderson. This project was supported financially by the Adaptation Research Network for Marine Biodiversity and Resources (NCCARF) and NCMCRS (University of Tasmania).
LITERATURE CITED
Aarssen LW, Clauss MJ (1992) Genotypic variation in fecun- dity allocation in Arabidopsis thaliana. J Ecol 80: 109−114
ABARES (Australian Bureau of Agricultural and Resource Economics and Sciences) (2011) Australian Fisheries Sta- tistics 2010. ABARES, Canberra
Andersen RA (2005) Algal culturing techniques. Elsevier Academic Press, London
Anderson MJ (2001) A new method for non-parametric mul- tivariate analysis of variance. Austral Ecol 26: 32−46
Anderson RJ, Levitt GJ, Share A (1996) Experimental inves- tigations for the mariculture of Gracilaria in Saldanha Bay, South Africa. J Appl Phycol 8: 421−430
Anderson MJ, Gorley RN, Clarke KR (2008) PERMANOVA+ for PRIMER: guide to software and statistical methods. Primer-e, Plymouth
Bartsch I, Wiencke C, Laepple T (2012) Global seaweed bio- geography under a changing climate: the prospected effects of temperature. In: Wiencke C, Bischof K (eds) Seaweed biology. Ecol Stud 219. Springer-Verlag, Berlin, p 383−406
Beaugrand G, Reid PC, Ibañez F, Lindley JA, Edwards M (2002) Reorganization of North Atlantic marine copepod biodiversity and climate. Science 296: 1692−1694
Benedetti-Cecchi L (2001) Variability in abundance of algae and invertebrates at different spatial scales on rocky sea shores. Mar Ecol Prog Ser 215: 79−92
Bolton JJ (1996) Patterns of species diversity and endemism in comparable temperate brown algal floras. Hydrobiolo- gia 326−327: 173−178
Bolton JJ, Levitt GJ (1985) Light and temperature require- ments for growth and reproduction in gametophytes of Ecklonia maxima (Alariaceae: Lamina riales). Mar Biol 87: 131−135
Cai W (2006) Antarctic ozone depletion causes an intensifi- cation of the Southern Ocean super-gyre circulation. Geo phys Res Lett 33: L03712, doi:10.1029/2005 GL024911
Cai W, Shi G, Cowan T, Bi D, Ribbe J (2005) The response of the Southern Annular Mode, the East Australian Cur- rent, and the southern mid-latitude ocean circulation to global warming. Geophys Res Lett 32: 94−97
Chapman ARO (1986) Age versus stage: an analysis of age- and size-specific mortality and reproduction in a popula- tion of Laminaria longicruris Pyl. J Exp Mar Biol Ecol 97: 113−122
Chesson P (1996) Matters of scale in the dynamics of popu- lations and communities. In: Floyd RB, Sheppard AW, Barro PJD (eds) Frontiers of population ecology. CSIRO, Melbourne, p 353−368
Connell SD, Russell BD, Turner DJ, Shepherd SA and others (2008) Recovering a lost baseline: missing kelp forests from a metropolitan coast. Mar Ecol Prog Ser 360: 63−72
Dayton PK (1985) Ecology of kelp communities. Annu Rev Ecol Syst 16: 215−245
Dayton PK, Tegner MJ, Edwards PB, Riser KL (1999) Tempo- ral and spatial scales of kelp demography: the role of oceanographic climate. Ecol Monogr 69: 219−250
Dean TA, Jacobsen FR (1986) Nutrient-limited growth of juvenile kelp, Macrocystis pyrifera, during the 1982− 1984 “El Niño” in southern California. Mar Biol 90: 597−601
Dolman G, Coleman MA (2008) Characterisation of micro - satellite loci in the habitat-forming kelp, Ecklonia radiata (Phaeophyceae, Laminariales). Conserv Genet 10: 657−660
Edyvane K (2003) Conservation, monitoring and recovery of threatened giant kelp (Macrocystis pyrifera) beds in Tas- mania—final report. Department of Primary Industries, Water and Environment, Hobart
Fowler-Walker MJ, Connell SD, Gillanders BM (2005) Vari- ation at local scales need not impede tests for broader scale patterns. Mar Biol 147: 823−831
Fowler-Walker MJ, Wernberg T, Connell SD (2006) Differ- ences in kelp morphology between wave sheltered and exposed localities: morphologically plastic or fixed traits? Mar Biol 148: 755−767
Gower JC (1967) A comparison of some methods of cluster analysis. Biometrics 23: 623−637
Graham MH, Vasquez JA, Buschmann AH (2007) Global ecology of the giant kelp Macrocystis: from ecotypes to ecosystems. Oceanogr Mar Biol Annu Rev 45: 39−88
Greenslade DJM (2001) The assimilation of ERS-2 signifi- cant wave height data in the Australian region. J Mar Syst 28: 141−160
Guiler ER (1960) The intertidal zone-forming species on rocky shores of the east Australian coast. J Ecol 48: 1−28
Hasselmann K (1988) PIPs and POPs: the reduction of com- plex dynamical systems using principal interaction and oscillation patterns. J Geophys Res 93: 11015−11021
Mar Ecol Prog Ser 483: 117–131, 2013
Hatcher BG, Kirkman H, Wood WF (1987) Growth of the kelp Ecklonia radiata near the northern limit of its range in western Australia. Mar Biol 95: 63−73
Holbrook NJ, Davidson J, Feng M, Hobday AJ, Lough JM, McGregor S, Risbey JS (2009) El Niño − Southern Oscil- lation. In: Poloczanska ES, Hobday AJ, Richardson AJ (eds) A marine climate change impacts and adaptation report card for Australia 2009. NCCARF, Gold Coast, p 1−25
Hughes MG, Heap AD (2010) National-scale wave energy resource assessment for Australia. Renew Energy 35: 1783−1791
Hurd CL, Hepburn CD, Currie KI, Raven JA, Hunter KA (2009) Testing the effects of ocean acidification on algal metabolism: considerations for experimental designs. J Phycol 45: 1236−1251
IPCC (Intergovernmental Panel on Climate Change) (2007) Summary for policymakers. In: Solomon S, Qin D, Man- ning M, Chen Z and others (eds) Contribution of Work- ing Group I to the Fourth Assessment Report of the Inter- governmental Panel on Climate Change. Cambridge University Press, Cambridge
Johnson C, Ling S, Ross J, Shepherd S, Miller K (2005) Establishment of the long-spined sea urchin (Centro - stephanus rodgersii) in Tasmania: first assessment of potential threats to fisheries. Final report. Fisheries Research and Development Corporation. University of Tasmania, Hobart
Johnson CR, Banks SC, Barrett NS, Cazassus F and others (2011) Climate change cascades: shifts in oceanography, species’ ranges and subtidal marine community dynam- ics in eastern Tasmania. J Exp Mar Biol Ecol 400: 17−32
Joska MAP, Bolton JJ (1987) In situ measurement of zoo - spore release and seasonality of reproduction in Ecklonia maxima (Alariaceae, Laminariales). Br Phycol J 22: 209−214
Kerswell AP (2006) Global biodiversity patterns of benthic marine algae. Ecology 87: 2479−2488
Kinlan BP, Graham MH, Sala E, Dayton PK (2003) Arrested development of giant kelp (Macrocystis pyrifera, Phaeo- phyceae) embryonic sporophytes: a mechanism for de - layed recruitment in perennial kelps? J Phycol 39: 47−57
Kirkman H (1981) The first year in the life history and the sur- vival of the juvenile marine macrophyte, Ecklonia radiata (Turn.) J. Agardh. J Exp Mar Biol Ecol 55: 243−254
Kirkman H (1984) Standing stock and production of Ecklo- nia radiata (C. Ag.) J. Agardh. J Exp Mar Biol Ecol 76: 119−130
Kirkman H, Kendrick GA (1997) Ecological significance and commercial harvesting of drifting and beach-cast macro- algae and seagrasses in Australia: a review. J Appl Phy- col 9: 311−326
Lewin J (1966) Silicon metabolism in diatoms v. germanium dioxide, a specific inhibitor of diatom growth. Phycologia 6: 1−12
Ling SD, Johnson CR, Frusher SD, Ridgway KR (2009a) Overfishing reduces resilience of kelp beds to climate- driven catastrophic phase shift. Proc Natl Acad Sci USA 106: 22341−22345
Ling SD, Johnson CR, Ridgway K, Hobday J, Haddon M (2009b) Climate-driven range extension of a sea urchin: inferring future trends by analysis of recent population dynamics. Glob Change Biol 15: 719−731
Mann KH (1973) Seaweeds: their productivity and strategy for growth. Science 182: 975−981
McLeod DJ, Hallegraeff GM, Hosie GW, Richardson AJ (2012) Climate-driven range expansion of the red-tide dinoflagellate Noctiluca scintillans into the Southern Ocean. J Plankton Res 34: 332−337
McLeod DJ, Hobday AJ, Lyle JM, Welsford DC (2012) A prey-related shift in the abundance of small pelagic fish in eastern Tasmania? ICES J Mar Sci 69: 953−960
Mohring MB, Wernberg T, Kendrick GA, Rule MJ (2013) Reproductive synchrony in a habitat-forming kelp and its relationship with environmental conditions. Mar Biol 160:119–126
Muñoz V, Hernández-Gonzáles MC, Buschmann AH, Gra- ham MH, Vásquez JA (2004) Variability in per capita oogonia and sporophyte production from giant kelp gametophytes (Macrocystis pyrifera, Phaeophyceae). Rev Chil Hist Nat 77: 639−647
Novaczek I (1980) The development and phenology of Eck- lonia radiata (C.Ag.) J. Ag. PhD thesis, University of Aukland
Novaczek I (1981) Stipe growth rings in Ecklonia radiata (C.Ag.) J.Ag. (Laminariales). Br Phycol J 16: 363−371
Novaczek I (1984a) Response of gametophytes of Ecklonia radiata (Laminariales) to temperature in saturating light. Mar Biol 82: 241−245
Novaczek I (1984b) Development and phenology of Ecklo- nia radiata at two depths in Goat Island Bay, New Zea - land. Mar Biol 81: 189−197
Parmesan C (2006) Ecological and evolutionary responses to recent climate change. Annu Rev Ecol Evol Syst 37: 637−669
Phillips JA (2001) Marine macroalgal biodiversity hotspots: Why is there high species richness and endemism in southern Australian marine benthic flora? Biodivers Con- serv 10: 1555−1577
Poloczanska ES, Babcock RC, Butler A, Hobday AJ and oth- ers (2007) Climate change and Australian marine life. Oceanogr Mar Biol Annu Rev 45: 407−478
Reed DC, Schroeter SC, Raimondi PT (2004) Spore supply and habitat availability as sources of recruitment limita- tion in the giant kelp Macrocystis pyrifera (Phaeo- phyceae). J Phycol 40: 275−284
Ridgway KR (2007) Long-term trend and decadal variability of the southward penetration of the East Australian Current. Geophys Res Lett 34: L13613, doi:10.1029/2007 GL030393
Ridgway K, Hill K (2009) The East Australian Current. In: Poloczanska ES, Hobday AJ, Richardson AJ (eds) A marine climate change impacts and adaptation report card for Australia. NCCARF, Gold Coast, p 1−16
Rochford DJ (1984) Nitrates in eastern Australian coastal waters. Aust J Mar Freshw Res 35: 385−397
Roleda MY, Morris JN, McGraw CM, Hurd CL (2012) Ocean acidification and seaweed reproduction: increased CO2
ameliorates the negative effect of lowered pH on meiospore germination in the giant kelp Macrocystis pyrifera (Laminariales, Phaeophyceae). Glob Change Biol 18: 854−864
Russell BD, Harley CDG, Wernberg T, Mieszkowska N, Widdicombe S, Hall-Spencer JM, Connell SD (2011) Pre- dicting ecosystem shifts requires new approaches that integrate the effects of climate change across entire sys- tems. Biol Lett 8: 164−166
Sagarin RD, Barry JP, Gilman SE, Baxter CH (1999) Climate- related change in an intertidal community over short and long time scales. Ecol Monogr 69: 465−490
130
Sanderson J (1990) Subtidal macroalgal studies in east and south eastern Tasmanian coastal waters. MS thesis, Uni- versity of Tasmania, Hobart
Serisawa Y, Yokohama Y, Aruga Y, Tanaka J (2002) Growth of Ecklonia cava (Laminariales, Phaeophyta) sporo- phytes transplanted to a locality with different tempera- ture conditions. Phycol Res 50: 201−207
Shea R, Chopin T (2007) Effects of germanium dioxide, an inhibitor of diatom growth, on the microscopic laboratory cultivation stage of the kelp, Laminaria saccharina. J Appl Phycol 19: 27−32
Smale DA, Wernberg T (2009) Satellite-derived SST data as a proxy for water temperature in nearshore benthic eco - lo gy. Mar Ecol Prog Ser 387: 27−37
Smith SDA, Simpson RD, Cairns SC (1996) The macrofau- nal community of Ecklonia radiata holdfasts: descrip- tion of the faunal assemblage and variation associated with differences in holdfast volume. Austral Ecol 21: 81−95
Staehr PA, Wernberg T (2009) Physiological responses of Ecklonia radiata (Laminariales) to a latitudinal gradient in ocean temperature. J Phycol 45: 91−99
Steneck RS, Graham MH, Bourque BJ, Corbett D, Erlandson JM, Estes JA, Tegner MJ (2002) Kelp forest ecosystems: biodiversity, stability, resilience and future. Environ Conserv 29: 436−459
tom Dieck I (1993) Temperature tolerance and survival in darkness of kelp gametophytes (Laminariales, Phaeo- phyta): ecological and biogeographical implications. Mar Ecol Prog Ser 100: 253−264
Underwood AJ, Chapman MG, Connell SD (2000) Observa- tions in ecology: you can’t make progress on processes without understanding the patterns. J Exp Mar Biol Ecol 250: 97−115
Van Patten MS, Yarish C (1993) Allocation of blade surface to reproduction in Laminaria longicruris of Long Island Sound (USA). Hydrobiologia 260−261: 173−181
van Tussenbroek BI (1989) Seasonal growth and composi- tion of fronds of Macrocystis pyrifera in the Falkland Islands. Mar Biol 100: 419−430
Wernberg T, Thomsen M (2005) The effect of wave exposure on the morphology of Ecklonia radiata. Aquat Bot 83: 61−70
Wernberg T, Coleman M, Fairhead A, Miller S, Thomsen M (2003) Morphology of Ecklonia radiata (Phaeophyta: Laminarales) along its geographic distribution in south- western Australia and Australasia. Mar Biol 143: 47−55
Wernberg T, Campbell A, Coleman MA, Connell SD and others (2009) Macroalgae and temperate rocky reefs. In: Poloczanska ES, Hobday AJ, Richardson AJ (eds) A marine climate change impacts and adaptation report card for Australia 2009. NCCARF, Gold Coast p 1−21
Wernberg T, Thomsen MS, Tuya F, Kendrick GA, Staehr PA, Toohey BD (2010) Decreasing resilience of kelp beds along a latitudinal temperature gradient: potential impli- cations for a warmer future. Ecol Lett 13: 685−694
Wernberg T, Russell BD, Moore PJ, Ling SD and others (2011) Impacts of climate change in a global hotspot for temperate marine biodiversity and ocean warming. J Exp Mar Biol Ecol 400: 7−16
Womersley HBS (1967) A critical survey of the marine algae of southern Australia. II. Phaeophyta. Aust J Bot 15: 189−270
Womersley HBS (1981) Biogeography of the Australasian marine macroalgae. In: Clayton MN, Kings RJ (eds) Marine botany: an Australian perspective. Longman Cheshire, Melbourne, p 292−307
Womersley HBS (1987) The marine benthic flora of Southern Australia. Part II. South Australian Government Printing Division, Adelaide
Wright LD (1976) Nearshore wave-power dissipation and the coastal energy regime of the Sydney-Jervis Bay region, New South Wales: a comparison. Aust J Mar Freshw Res 27: 633−640
Wright JT, de Nys R, Poore AGB, Steinberg PD (2004) Chemical defense in a marine alga: heritability and the potential for selection by herbivores. Ecology 85: 2946−2959
Young IR, Zieger S, Babanin AV (2011) Global trends in wind speed and wave height. Science 332:451–455
Editorial responsibility: Laura Airoldi, Ravenna, Italy
Submitted: July 23, 2012; Accepted: January 11, 2013 Proofs received from author(s): April 30, 2013
http://dx.doi.org/10.1890/03-4041
http://dx.doi.org/10.1071/MF9760633
http://dx.doi.org/10.1071/BT9670189
http://dx.doi.org/10.1016/j.jembe.2011.02.021
http://dx.doi.org/10.1111/j.1461-0248.2010.01466.x
http://dx.doi.org/10.1007/s00227-003-1069-9
http://dx.doi.org/10.1016/j.aquabot.2005.05.007
http://dx.doi.org/10.1007/BF00391158
http://dx.doi.org/10.1016/S0022-0981(00)00181-7
http://dx.doi.org/10.3354/meps100253
http://dx.doi.org/10.1017/S0376892902000322
http://dx.doi.org/10.1111/j.1529-8817.2008.00635.x
http://dx.doi.org/10.1111/j.1442-9993.1996.tb00587.x
http://dx.doi.org/10.3354/meps08132
http://dx.doi.org/10.1007/s10811-006-9107-x
cite10:
cite12:
cite14:
cite21:
cite23:
cite16:
cite30:
cite25:
cite18:
cite32:
cite41:
cite43:
cite36:
cite8:
cite38:
cite50:
cite45:
cite54:
cite47:
cite29:
cite56:
cite49:
cite63:
cite58:
cite65:
cite70:
cite67:
cite69:
cite5:
cite9:
cite11:
cite13:
cite20:
cite22:
cite15:
cite24:
cite17:
cite2:
cite26:
cite40:
cite19:
cite28:
cite42:
cite6:
cite51:
cite35:
cite33:
cite39:
cite53:
cite31:
cite55:
cite48:
cite44:
cite57:
cite60:
cite62:
cite59:
cite64:
cite71:
cite68:
cite3:
cite7: